首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 359 毫秒
1.
2.
3.
Protein kinase D (PKD) is a serine/threonine protein kinase rapidly activated by G protein-coupled receptor (GPCR) agonists via a protein kinase C (PKC)-dependent pathway. Recently, PKD has been implicated in the regulation of long term cellular activities, but little is known about the mechanism(s) of sustained PKD activation. Here, we show that cell treatment with the preferential PKC inhibitors GF 109203X or Gö 6983 blocked rapid (1–5-min) PKD activation induced by bombesin stimulation, but this inhibition was greatly diminished at later times of bombesin stimulation (e.g. 45 min). These results imply that GPCR-induced PKD activation is mediated by early PKC-dependent and late PKC-independent mechanisms. Western blot analysis with site-specific antibodies that detect the phosphorylated state of the activation loop residues Ser744 and Ser748 revealed striking PKC-independent phosphorylation of Ser748 as well as Ser744 phosphorylation that remained predominantly but not completely PKC-dependent at later times of bombesin or vasopressin stimulation (20–90 min). To determine the mechanisms involved, we examined activation loop phosphorylation in a set of PKD mutants, including kinase-deficient, constitutively activated, and PKD forms in which the activation loop residues were substituted for alanine. Our results show that PKC-dependent phosphorylation of the activation loop Ser744 and Ser748 is the primary mechanism involved in early phase PKD activation, whereas PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies identify a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.A rapid increase in the synthesis of lipid-derived second messengers with subsequent activation of protein phosphorylation cascades has emerged as a fundamental signal transduction mechanism triggered by multiple extracellular stimuli, including hormones, neurotransmitters, chemokines, and growth factors (1). Many of these agonists bind to G protein-coupled receptors (GPCRs),4 activate heterotrimeric G proteins and stimulate isoforms of the phospholipase C family, including β, γ, δ, and ε (reviewed in Refs. 1 and 2). Activated phospholipase Cs catalyze the hydrolysis of phosphatidylinositol 4,5-bisphosphate to produce the second messengers inositol 1,4,5-trisphosphate and diacylglycerol (DAG). Inositol 1,4,5-trisphosphate mobilizes Ca2+ from intracellular stores (3, 4) whereas DAG directly activates the classic (α, β, and γ) and novel (δ, ε, η, and θ) isoforms of PKC (57). Although it is increasingly recognized that each PKC isozyme has specific functions in vivo (58), the mechanisms by which PKC-mediated signals are propagated to critical downstream targets remain incompletely defined.PKD, also known initially as PKCμ (9, 10), and two recently identified serine protein kinases termed PKD2 (11) and PKCν/PKD3 (12, 13), which are similar in overall structure and primary amino acid sequence to PKD (14), constitute a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase group (15) and separate from the previously identified PKCs (14). Salient features of PKD structure include an N-terminal regulatory region containing a tandem repeat of cysteine-rich zinc finger-like motifs (termed the cysteine-rich domain) that confers high affinity binding to phorbol esters and DAG (9, 16, 17), followed by a pleckstrin homology (PH) domain that negatively regulates catalytic activity (18, 19). The C-terminal region of the PKDs contains its catalytic domain, which is distantly related to Ca2+-regulated kinases.In unstimulated cells, PKD is in a state of low kinase catalytic activity maintained by the N-terminal domain, which represses the catalytic activity of the enzyme by autoinhibition. Consistent with this model, deletions or single amino acid substitutions in the PH domain result in constitutive kinase activity (1820). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (21). In response to cellular stimuli, PKD is converted from a low activity form into a persistently active form that is retained during isolation from cells, as shown by in vitro kinase assays performed in the absence of lipid co-activators (21, 22). PKD activation has been demonstrated in response to engagement of specific GPCRs either by regulatory peptides (2330) or lysophosphatidic acid (27, 31, 32); signaling through Gq, G12, Gi, and Rho (27, 3134); activation of receptor tyrosine kinases, such as the platelet-derived growth factor receptor (23, 35, 36); cross-linking of B-cell receptor and T-cell receptor in B and T lymphocytes, respectively (3740); and oxidative stress (4144).Throughout these studies, multiple lines of evidence indicated that PKC activity is necessary for rapid PKD activation within intact cells. For example, rapid PKD activation was selectively and potently blocked by cell treatment with preferential PKC inhibitors (e.g. GF 109203X or Gö 6983) that do not directly inhibit PKD catalytic activity (21, 22), implying that PKD activation in intact cells is mediated, directly or indirectly, through PKCs. In line with this conclusion, cotransfection of PKD with active mutant forms of “novel” PKCs (PKCs δ, ε, η, and θ) resulted in robust PKD activation in the absence of cell stimulation (21, 4446). Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade in response to multiple GPCR agonists in a broad range of cell types, including normal and cancer cells (reviewed in Ref. 14). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as the activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (reviewed in Ref. 14). Collectively, these findings demonstrated the existence of rapidly activated PKC-PKD protein kinase cascade(s) and raised the possibility that some PKC-dependent biological responses involve PKD acting as a downstream effector.PKD has been reported recently to mediate several important cellular activities and processes, including signal transduction (30, 4749), chromatin modification (50), Golgi organization and function (51, 52), c-Jun function (47, 53, 54), NFκB-mediated gene expression (43, 55, 56), and cell survival, migration, and differentiation and DNA synthesis and proliferation (reviewed in Ref. 14). Thus, mounting evidence indicates that PKD has a remarkable diversity of both its signal generation and distribution and its potential for complex regulatory interactions with multiple downstream pathways, leading to multiple responses, including long term cellular events. Despite increasing recognition of its importance, very little is known about the mechanism(s) of sustained PKD activation as opposed to the well documented rapid, PKC-dependent PKD activation.The results presented here demonstrate that prolonged GPCR-induced PKD activation is mediated by sequential PKC-dependent and PKC-independent phases of regulation. We report here, for the first time, that PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies expand previous models of PKD regulation by identifying a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.  相似文献   

4.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  相似文献   

5.
6.
7.
Arrestins bind active phosphorylated G protein-coupled receptors, precluding G protein activation and channeling signaling to alternative pathways. Arrestins also function as mitogen-activated protein kinase (MAPK) scaffolds, bringing together three components of MAPK signaling modules. Here we have demonstrated that all four vertebrate arrestins interact with JNK3, MKK4, and ASK1, but only arrestin3 facilitates JNK3 activation. Thus, the functional specificity of arrestins is not determined by differential binding of the kinases. Using receptor binding-impaired mutant, we have shown that free arrestin3 readily promotes JNK3 phosphorylation. We identified key arrestin-binding elements in JNK3 and ASK1 and investigated the molecular interactions of arrestin2 and arrestin3 and their individual domains with the components of the two MAPK cascades, ASK1-MKK4-JNK3 and c-Raf-1-MEK1-ERK2. We found that both arrestin domains interact with all six kinases. These findings shed new light on the mechanism of arrestin-mediated MAPK activation and the spatial arrangement of the three kinases on arrestin molecule.Arrestins are multifunctional regulators of cell signaling (1, 2). Arrestins, which bind active phosphorylated G protein-coupled receptors (GPCRs),2 which play a major role in receptor desensitization and internalization (3, 4). With the identification of numerous non-receptor binding partners, the classical paradigm of arrestin function has been expanded, implicating arrestins in mitogen-activated protein kinase (MAPK) activation, protein ubiquitination, chemotaxis, apoptosis, and other cellular functions (2, 5-11).The first indication that arrestins function as signaling adapters came from the studies of arrestin-dependent c-Src recruitment to the receptors, which results in the activation of extracellular signal-regulated kinases (ERK1/2) (10, 12, 13). Subsequently, arrestin2 and arrestin3 in complex with different receptors were reported to scaffold JNK3 (9), ERK1/2 (8, 14), and p38 (15, 16) activation cascades. Although arrestins play an important role in regulating different MAPK pathways, the mechanism of arrestin-dependent assembly of MAP kinases into a signaling complex remains largely unexplored. Existing models have limited predictive value. For example, the idea that JNK3 is activated solely by arrestin3 because this arrestin subtype has unique ability to bind JNK3 (9, 17) was not supported by further experimentation (18-20). Similarly, the hypothesis that only receptor-bound arrestins interact with MAP kinases (8, 9) was not confirmed (17-20).Here we addressed several key mechanistic issues in arrestin-dependent MAPK signaling. First, we show that the scaffolding function is not limited to receptor-bound arrestin; free arrestin3 facilitates ASK1-mediated JNK3 activation, indicating that arrestins are not exclusively receptor-regulated adapters as thought previously. Second, we show that all four mammalian arrestins bind each component of the JNK3 cascade with comparable affinity, demonstrating that binding does not necessarily translate into activation. This finding establishes the mechanistic basis of the “dominant-negative” effect of certain arrestin subtypes. Third, using truncated forms of ASK1 and JNK3, we identified the major arrestin-binding elements of these two kinases. Finally, we show that every kinase in JNK3 and ERK2 activation cascades binds both arrestin domains. Based on these findings, we propose a functional model of arrestin-dependent regulation of MAPK activity and a new structural model of the arrestin-MAPK multiprotein signaling complex.  相似文献   

8.
9.
10.
Mutations in SHP-2 phosphatase (PTPN11) that cause hyperactivation of its catalytic activity have been identified in Noonan syndrome and various childhood leukemias. Recent studies suggest that the gain-of-function (GOF) mutations of SHP-2 play a causal role in the pathogenesis of these diseases. However, the molecular mechanisms by which GOF mutations of SHP-2 induce these phenotypes are not fully understood. Here, we show that GOF mutations in SHP-2, such as E76K and D61G, drastically increase spreading and migration of various cell types, including hematopoietic cells, endothelial cells, and fibroblasts. More importantly, in vivo angiogenesis in SHP-2 D61G knock-in mice is also enhanced. Mechanistic studies suggest that the increased cell migration is attributed to the enhanced β1 integrin outside-in signaling. In response to β1 integrin cross-linking or fibronectin stimulation, activation of ERK and Akt kinases is greatly increased by SHP-2 GOF mutations. Also, integrin-induced activation of RhoA and Rac1 GTPases is elevated. Interestingly, mutant cells with the SHP-2 GOF mutation (D61G) are more sensitive than wild-type cells to the suppression of cell motility by inhibition of these pathways. Collectively, these studies reaffirm the positive role of SHP-2 phosphatase in cell motility and suggest a new mechanism by which SHP-2 GOF mutations contribute to diseases.SHP-2, a multifunctional SH2 domain-containing protein-tyrosine phosphatase implicated in diverse cell signaling processes (13), plays a critical role in cellular function. Homozygous deletion of Exon 2 (4) or Exon 3 (5) of the SHP-2 gene (PTPN11) in mice leads to early embryonic lethality prior to and at midgestation, respectively. SHP-2 null mutant mice die much earlier, at peri-implantation (4). Exon 3 deletion mutation of SHP-2 blocks hematopoietic potential of embryonic stem cells both in vitro and in vivo (68), whereas SHP-2 null mutation causes inner cell mass death and diminished trophoblast stem cell survival (4). Recent studies on SHP-2 conditional knock-out or tissue-specific knock-out mice have further revealed an array of important functions of this phosphatase in various physiological processes (912). The phenotypes demonstrated by loss of SHP-2 function are apparently attributed to the role of SHP-2 in the cell signaling pathways induced by growth factors/cytokines. SHP-2 generally promotes signal transmission in growth factor/cytokine signaling in both catalytic-dependent and -independent fashion (13). The positive role of SHP-2 in the intracellular signaling processes, in particular, the ERK3 and PI3K/Akt kinase pathways, has been well established, although the underlying mechanism remains elusive, in particular, the signaling function of the catalytic activity of SHP-2 in these pathways is poorly understood.In addition to the role of SHP-2 in cell proliferation and differentiation, the phenotypes induced by loss of SHP-2 function may be associated with its role in cell migration. Indeed, dominant negative SHP-2 disrupts Xenopus gastrulation, causing tail truncations (13, 14). Targeted Exon 3 deletion mutation in SHP-2 results in decreased cell spreading, migration (15, 16), and impaired limb development in the chimeric mice (7). The role of SHP-2 in cell adhesion and migration has also been demonstrated by catalytically inactive mutant SHP-2-overexpressing cells (1720). The molecular mechanisms by which SHP-2 regulates these cellular processes, however, have not been well defined. For example, the role of SHP-2 in the activation of the Rho family small GTPases that is critical for cell motility is still controversial. Both positive (19, 21, 22) and negative roles (18, 23) for SHP-2 in this context have been reported. Part of the reason for this discrepancy might be due to the difference in the cell models used. Catalytically inactive mutant SHP-2 was often used to determine the role of SHP-2 in cell signaling. In the catalytically inactive mutant SHP-2-overexpressing cells, the catalytic activity of endogenous SHP-2 is inhibited. However, as SHP-2 also functions independent of its catalytic activity, overexpression of catalytically deficient SHP-2 may also increase its scaffolding function, generating complex effects.The critical role of SHP-2 in cellular function is further underscored by the identification of SHP-2 mutations in human diseases. Genetic lesions in PTPN11 that cause hyperactivation of SHP-2 catalytic activity have been identified in the developmental disorder Noonan syndrome (24) and various childhood leukemias, including juvenile myelomonocytic leukemia (JMML), B cell acute lymphoblastic leukemia, and acute myeloid leukemia (25, 26). In addition, activating mutations in SHP-2 have been identified in sporadic solid tumors (27). The SHP-2 mutations appear to play a causal role in the development of these diseases as SHP-2 mutations and other JMML-associated Ras or Neurofibromatosis 1 mutations are mutually exclusive in the patients (2427). Moreover, single SHP-2 gain-of-function (GOF) mutations are sufficient to induce Noonan syndrome, cytokine hypersensitivity in hematopoietic progenitor cells, and JMML-like myeloproliferative disease in mice (2832). Gain-of-function cell models derived from the newly available SHP-2 GOF mutation (D61G) knock-in mice (28) now provide us with a good opportunity to clarify the role of SHP-2 in cell motility. Unlike the dominant negative approach in which overexpression of mutant forms of SHP-2 generates complex effects, the SHP-2 D61G knock-in model eliminates this possibility as the mutant SHP-2 is expressed at the physiological level (28). Additionally, defining signaling functions of GOF mutant SHP-2 in cell movement can also help elucidate the molecular mechanisms by which SHP-2 mutations contribute to the relevant diseases.  相似文献   

11.
The Src homology phosphotyrosyl phosphatase 2 (SHP2) plays a positive role in HER2-induced signaling and transformation, but its mechanism of action is poorly understood. Given the significance of HER2 in breast cancer, defining a mechanism for SHP2 in the HER2 signaling pathway is of paramount importance. In the current report we show that SHP2 positively modulates the Ras-extracellular signal-regulated kinase 1 and 2 and the phospoinositide-3-kinase-Akt pathways downstream of HER2 by increasing the half-life the activated form of Ras. This is accomplished by dephosphorylating an autophosphorylation site on HER2 that serves as a docking platform for the SH2 domains of the Ras GTPase-activating protein (RasGAP). The net effect is an increase in the intensity and duration of GTP-Ras levels with the overall impact of enhanced HER2 signaling and cell transformation. In conformity to these findings, the HER2 mutant that lacks the SHP2 target site exhibits an enhanced signaling and cell transformation potential. Therefore, SHP2 promotes HER2-induced signaling and transformation at least in part by dephosphorylating a negative regulatory autophosphorylation site. These results suggest that SHP2 might serve as a therapeutic target against breast cancer and other cancers characterized by HER2 overexpression.The Src homology phosphotyrosyl phosphatase 2 (SHP2)2 functions as a positive effector of cell growth and survival (14), migration and invasion (58), and morphogenesis and transformation (911). In receptor-tyrosine kinase signaling (1214), SHP2 positively transduces the Ras-extracellular signal-regulated kinase 1 and 2 (ERK1/2) and the phosphoinositide-3-kinase-Akt (or protein kinase B) signaling pathways. SHP2 also promotes cell transformation induced by the constitutively active form of fibroblast growth factor receptor 3 and v-Src (9, 11). The discovery of germline-activating SHP2 mutations in Noonan and LEOPARD syndrome patients (1518) and the subsequent experimental demonstration of these phenotypes in knockin and transgenic mice expressing these mutants (19, 20) has led to the conclusion that disregulation of SHP2 is responsible for these disease states. Furthermore, somatic activating SHP2 mutations were discovered in juvenile myelomonocytic leukemia, acute myelogenous leukemia, and chronic myelomonocytic (18, 21) and are suggested to play a causative role.SHP2 possesses two Src homology 2 (SH2) domains in the N-terminal region that allow the protein to localize to substrate microdomains after tyrosyl phosphorylation of interacting proteins. The phosphotyrosyl phosphatase (PTP) domain in the C-terminal region is responsible for dephosphorylation of target substrates (13, 22). Mutation of the critical Cys residue in the active site of SHP2 abolishes its phosphatase activity, leading to the production of a dominant-negative protein (23). The activity of SHP2 is regulated by an intramolecular conformational switch. SHP2 assumes a “closed conformation” when inactive and an “open conformation” when active. In the closed conformation the N-SH2 domain interacts with the PTP domain, physically impeding the activity of the enzyme. Upon engagement of the SH2 domains with phosphotyrosine, the PTP domain is relieved of autoinhibition and dephosphorylates target substrates (2326). Interaction between specific residues on the N-SH2 and the PTP domains mediates the closed conformation. Mutation of these residues leads to a constitutively active SHP2, and the occurrence of such mutations in humans causes the development of Noonan syndrome and associated leukemia (1618).Recently, we have shown that inhibition of SHP2 in the HER2-positive breast cancer cell lines abolishes mitogenic and cell survival signaling and reverses transformation, leading to differentiation of malignant cells into a normal breast epithelial phenotype (27). Given the significance of HER2 in breast cancer, the finding that SHP2 plays a positive role was very interesting. We, thus, sought to investigate the molecular mechanism that underlies the positive role of SHP2 in HER2-induced signaling and transformation. To do so, it was first necessary to decipher the identity of SHP2 substrates whose dephosphorylation promotes the oncogenic functions of HER2. Using the recently developed substrate-trapping mutant of SHP2 as a reagent (28), we have identified HER2 itself as an SHP2 substrate. We have further shown that SHP2 dephosphorylates an autophosphorylation site on HER2 that serves as a docking site for the SH2 domains of the Ras GTPase-activating protein (Ras-GAP), the down-regulator of Ras. This effect of SHP2 increases the intensity and duration of GTP-Ras levels with the overall impact of enhanced HER2 signaling and cell transformation.  相似文献   

12.
The presence of extensive reciprocal conformational freedom between the catalytic and the hemopexin-like domains of full-length matrix metalloproteinase-1 (MMP-1) is demonstrated by NMR and small angle x-ray scattering experiments. This finding is discussed in relation to the essentiality of the hemopexin-like domain for the collagenolytic activity of MMP-1. The conformational freedom experienced by the present system, having the shortest linker between the two domains, when compared with similar findings on MMP-12 and MMP-9 having longer and the longest linker within the family, respectively, suggests this type of conformational freedom to be a general property of all MMPs.Matrix metalloproteinases (MMP)2 are extracellular hydrolytic enzymes involved in a variety of processes including connective tissue cleavage and remodeling (13). All 23 members of the family are able to cleave simple peptides derived from connective tissue components such as collagen, gelatin, elastin, etc. A subset of MMPs is able to hydrolyze more resistant polymeric substrates, such as cross-linked elastin, and partially degraded collagen forms, such as gelatin and type IV collagens (4). Intact triple helical type I–III collagen is only attacked by collagenases MMP-1, MMP-8, and MMP-13 and by MMP-2 and MMP-14 (512). Although the detailed mechanism of cleavage of single chain peptides by MMP has been largely elucidated (1319), little is known about the process of hydrolysis of triple helical collagen. In fact, triple helical collagen cannot be accommodated in the substrate-binding groove of the catalytic site of MMPs (9).All MMPs (but MMP-7) in their active form are constituted by a catalytic domain (CAT) and a hemopexin-like domain (HPX) (2022). The CAT domain contains two zinc ions and one to three calcium ions. One zinc ion is at the catalytic site and is responsible for the activity, whereas the other metal ions have structural roles. The isolated CAT domains retain full catalytic activity toward simple peptides and single chain polymeric substrates such as elastin, whereas hydrolysis of triple helical collagen also requires the presence of the HPX domain (9, 2325). It has been shown that the isolated CAT domain regains a small fraction of the activity of the full-length (FL) protein when high amounts of either inactivated full-length proteins or isolated HPX domains are added to the assay solution (9). Finally, it has been shown that the presence of the HPX domain alone alters the CD spectrum of triple helical collagen in a way that suggests its partial unwinding (26, 27). It is tempting to speculate that full-length collagenases attack collagen by first locally unwinding the triple helical structure with the help of the HPX domain and then cleaving the resulting, exposed, single filaments (9, 28).Until 2007, three-dimensional structures of full-length MMPs had been reported only for collagenase MMP-1 (2931) and gelatinase MMP-2 (32). The structures of the two proteins are very similar and show a compact arrangement of the two domains, which are connected by a short linker (14 and 20 amino acids, respectively). It is difficult to envisage that rigid and compact molecules of this type can interact with triple helical collagen in a way that can lead to first unwinding and then cleavage of individual filaments. It has been recently suggested that such concerted action could occur much more easily if the two domains could enjoy at least a partial conformational independence (9). Slight differences in the reciprocal orientation of the CAT and HPX domains of MMP-1 in the presence (29) and absence (30, 31) of the prodomain were indeed taken as a hint that the two domains could experience relative mobility (29).Two recent solution studies have shown that conformational independence is indeed occurring in gelatinase MMP-9 (33) and elastase MMP-12 (34), whereas the x-ray structure of the latter (34) is only slightly less compact than those of MMP-1 (2931) and MMP-2 (32). Among MMPs, MMP-9 features an exceptionally long linker (68 amino acid) (33, 35), which in fact constitutes a small domain by itself (the O-glycosylated domain) (33), and therefore, this inspiring observation can hardly be taken as evidence that conformational freedom is a general characteristic of the two-domain MMPs. MMP-12 features a much more normal 16-amino acid linker, thereby making more probable a general functional role for this conformational freedom (34). However, both MMP-9 and MMP-12 retain their full catalytic activity against their substrates even when deprived of the HPX domain (9). Therefore, the question remains of whether conformational freedom is also a required characteristic for those MMPs that are only active as full-length proteins, i.e. collagenases. Interestingly, the three collagenases (MMP-1, MMP-8, and MMP-13) have the shortest linker (14 amino acids) among all MMPs. Demonstrating or negating the presence of conformational freedom in one of these collagenases would therefore constitute a significant step forward to formulate mechanistic hypotheses on their collagenolytic activity.Our recent studies on MMP-12 in solution (34) have shown that a combination of NMR relaxation studies and small angle x-ray scattering (SAXS) is enough to show the presence and the extent of the relative conformational freedom of the two domains of MMPs. Here we apply the same strategy to full-length MMP-1 and show that sizable conformational freedom is indeed experienced even by this prototypical collagenase, although somewhat less pronounced than that observed for MMP-12.  相似文献   

13.
mTORC1 contains multiple proteins and plays a central role in cell growth and metabolism. Raptor (regulatory-associated protein of mammalian target of rapamycin (mTOR)), a constitutively binding protein of mTORC1, is essential for mTORC1 activity and critical for the regulation of mTORC1 activity in response to insulin signaling and nutrient and energy sufficiency. Herein we demonstrate that mTOR phosphorylates raptor in vitro and in vivo. The phosphorylated residues were identified by using phosphopeptide mapping and mutagenesis. The phosphorylation of raptor is stimulated by insulin and inhibited by rapamycin. Importantly, the site-directed mutation of raptor at one phosphorylation site, Ser863, reduced mTORC1 activity both in vitro and in vivo. Moreover, the Ser863 mutant prevented small GTP-binding protein Rheb from enhancing the phosphorylation of S6 kinase (S6K) in cells. Therefore, our findings indicate that mTOR-mediated raptor phosphorylation plays an important role on activation of mTORC1.Mammalian target of rapamycin (mTOR)2 has been shown to function as a critical controller in cellular growth, survival, metabolism, and development (1). mTOR, a highly conserved Ser-Thr phosphatidylinositol 3-kinase-related protein kinase, structurally forms two distinct complexes, mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2), each of which catalyzes the phosphorylation of different substrates (1). The best characterized substrates for mTORC1 are eIF4E-binding protein (4E-BP, also known as PHAS) and p70 S6 kinase (S6K) (1), whereas mTORC2 phosphorylates the hydrophobic and turn motifs of protein kinase B (Akt/protein kinase B) (2) and protein kinase C (3, 4). mTORC1 constitutively consists of mTOR, raptor, and mLst8/GβL (1), whereas the proline-rich Akt substrate of 40 kDa (PRAS40) is a regulatory component of mTORC1 that disassociates after growth factor stimulation (5, 6). Raptor is essential for mTORC1 activity by providing a substrate binding function (7) but also plays a regulatory role on mTORC1 with stimuli of growth factors and nutrients (8). In response to insulin, raptor binding to substrates is elevated through the release of the competitive inhibitor PRAS40 from mTORC1 (9, 10) because PRAS40 and the substrates of mTORC1 (4E-BP and S6K) appear to bind raptor through a consensus sequence, the TOR signaling (TOS) motif (1014). In response to amino acid sufficiency, raptor directly interacts with a heterodimer of Rag GTPases and promotes mTORC1 localization to the Rheb-containing vesicular compartment (15).mTORC1 integrates signaling pathways from growth factors, nutrients, energy, and stress, all of which generally converge on the tuberous sclerosis complex (TSC1-TSC2) through the phosphorylation of TSC2 (1). Growth factors inhibit the GTPase-activating protein activity of TSC2 toward the small GTPase Rheb via the PI3K/Akt pathway (16, 17), whereas energy depletion activates TSC2 GTPase-activating protein activity by stimulating AMP-activated protein kinase (AMPK) (18). Rheb binds directly to mTOR, albeit with very low affinity (19), and upon charging with GTP, Rheb functions as an mTORC1 activator (6). mTORC1 complexes isolated from growth factor-stimulated cells show increased kinase activity yet do not contain detectable levels of associated Rheb. Therefore, how Rheb-GTP binding to mTOR leads to an increase in mTORC1 activity toward substrates, and what the role of raptor is in this activation is currently unknown. More recently, the AMPK and p90 ribosomal S6 kinase (RSK) have been reported to directly phosphorylate raptor and regulate mTORC1 activity. The phosphorylation of raptor directly by AMPK reduced mTORC1 activity, suggesting an alternative regulation mechanism independent of TSC2 in response to energy supply (20). RSK-mediated raptor phosphorylation enhances mTORC1 activity and provides a mechanism whereby stress may activate mTORC1 independent of the PI3K/Akt pathway (21). Therefore, the phosphorylation status of raptor can be critical for the regulation of mTORC1 activity.In this study, we investigated phosphorylation sites in raptor catalyzed by mTOR. Using two-dimensional phosphopeptide mapping, we found that Ser863 and Ser859 in raptor were phosphorylated by mTOR both in vivo and in vitro. mTORC1 activity in vitro and in vivo is associated with the phosphorylation of Ser863 in raptor.  相似文献   

14.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

15.
p53 is an important tumor suppressor regulating the cell cycle at multiple stages in higher vertebrates. The p53 gene is frequently deleted or mutated in human cancers, resulting in loss of p53 activity. This leads to centrosome amplification, aneuploidy, and tumorigenesis, three phenotypes also observed after overexpression of the oncogenic kinase Aurora A. Accordingly, recent studies have focused on the relationship between these two proteins. p53 and Aurora A have been reported to interact in mammalian cells, but the function of this interaction remains unclear. We recently reported that Xenopus p53 can inhibit Aurora A activity in vitro but only in the absence of TPX2. Here we investigate the interplay between Xenopus Aurora A, TPX2, and p53 and show that newly synthesized TPX2 is required for nearly all Aurora A activation and for full p53 synthesis and phosphorylation in vivo during oocyte maturation. In vitro, phosphorylation mediated by Aurora A targets serines 129 and 190 within the DNA binding domain of p53. Glutathione S-transferase pull-down studies indicate that the interaction occurs via the p53 transactivation domain and the Aurora A catalytic domain around the T-loop. Our studies suggest that targeting of TPX2 might be an effective strategy for specifically inhibiting the phosphorylation of Aurora A substrates, including p53.Aurora A is an oncogenic protein kinase that is active in mitosis and plays important roles in spindle assembly and centrosome function (1). Overexpression of either human or Xenopus Aurora A transforms mammalian cells, but only when the p53 pathway is altered (24). Aurora A is localized on centrosomes during mitosis, and overexpression of the protein leads to centrosome amplification and aneuploidy (2, 3, 5, 6), two likely contributors to genomic instability (7, 8). Because of its oncogenic potential and amplification in human tumors, considerable attention has been focused on the mechanism of Aurora A activation in mitosis. Evidence from several laboratories indicates that activation occurs as a result of phosphorylation of a threonine residue in the T-loop of the kinase (4, 9, 10). Purification of Aurora A-activating activity from M phase Xenopus egg extracts led to an apparent activation mechanism in which autophosphorylation at the T-loop is stimulated by binding of the targeting protein for Xklp2 (TPX2) (1114). On the other hand, it has been shown that Aurora A activity can be inhibited by interaction with several proteins, including PP1 (protein phosphatase 1), AIP (Aurora A kinase-interacting protein), and, more recently, p53 (9, 1517).p53 is a well known tumor suppressor able to drive cell cycle arrest, apoptosis, or senescence when DNA is damaged or cell integrity is threatened (18, 19). In human cancers, the p53 gene is frequently deleted or mutated, leading to inactivation of p53 functions (20). p53 protein is almost undetectable in “normal cells,” mainly due to its instability. Indeed, during a normal cell cycle, p53 associates with Mdm2 in the nucleus and thereafter undergoes nuclear exclusion, allowing its ubiquitination and subsequent degradation (21). In cells under stress, p53 is stabilized through the disruption of its interaction with Mdm2 (21), leading to p53 accumulation in the nucleus and triggering different responses, as described above.Although p53 has mostly been characterized as a nuclear protein, it has also been shown to localize on centrosomes (2224) and regulate centrosome duplication (23, 24). Centrosomes are believed to act as scaffolds that concentrate many regulatory molecules involved in signal transduction, including multiple protein kinases (25). Thus, centrosomal localization of p53 might be important for its own regulation by phosphorylation/dephosphorylation, and one of its regulators could be the mitotic kinase Aurora A. Indeed, phenotypes associated with the misexpression of these two proteins are very similar. For example, overexpression of Aurora A kinase leads to centrosome amplification, aneuploidy, and tumorigenesis, and the same effects are often observed after down-regulation of p53 transactivation activity or deletion/mutation of its gene (26, 27).Several recent studies performed in mammalian models show interplay between p53 and Aurora A, with each protein having the ability to inhibit the other, depending on the stage of the cell cycle and the stress level of the cell (17, 28, 29). These studies reported that p53 is a substrate of Aurora A, and serines 215 and 315 were demonstrated to be the two major Aurora A phosphorylation sites in human p53 in vitro and in vivo. Phosphorylation of Ser-215 within the DNA binding domain of human p53 inhibited both p53 DNA binding and transactivation activities (29). Recently, our group showed that Xenopus p53 is able to inhibit Aurora A kinase activity in vitro, but this inhibitory effect can be suppressed by prior binding of Aurora A to TPX2 (9). Contrary to somatic cells, where p53 is nuclear, unstable, and expressed at a very low level, p53 is highly expressed in the cytoplasm of Xenopus oocytes and stable until later stages of development (30, 31). The high concentration of both p53 and Aurora A in the oocyte provided a suitable basis for investigating p53-Aurora A interaction and also evaluating Xenopus p53 as a substrate of Aurora A.  相似文献   

16.
17.
18.
19.
Formin-homology (FH) 2 domains from formin proteins associate processively with the barbed ends of actin filaments through many rounds of actin subunit addition before dissociating completely. Interaction of the actin monomer-binding protein profilin with the FH1 domain speeds processive barbed end elongation by FH2 domains. In this study, we examined the energetic requirements for fast processive elongation. In contrast to previous proposals, direct microscopic observations of single molecules of the formin Bni1p from Saccharomyces cerevisiae labeled with quantum dots showed that profilin is not required for formin-mediated processive elongation of growing barbed ends. ATP-actin subunits polymerized by Bni1p and profilin release the γ-phosphate of ATP on average >2.5 min after becoming incorporated into filaments. Therefore, the release of γ-phosphate from actin does not drive processive elongation. We compared experimentally observed rates of processive elongation by a number of different FH2 domains to kinetic computer simulations and found that actin subunit addition alone likely provides the energy for fast processive elongation of filaments mediated by FH1FH2-formin and profilin. We also studied the role of FH2 structure in processive elongation. We found that the flexible linker joining the two halves of the FH2 dimer has a strong influence on dissociation of formins from barbed ends but only a weak effect on elongation rates. Because formins are most vulnerable to dissociation during translocation along the growing barbed end, we propose that the flexible linker influences the lifetime of this translocative state.Formins are multidomain proteins that assemble unbranched actin filament structures for diverse processes in eukaryotic cells (reviewed in Ref. 1). Formins stimulate nucleation of actin filaments and, in the presence of the actin monomer-binding protein profilin, speed elongation of the barbed ends of filaments (2-6). The ability of formins to influence elongation depends on the ability of single formin molecules to remain bound to a growing barbed end through multiple rounds of actin subunit addition (7, 8). To stay associated during subunit addition, a formin molecule must translocate processively on the barbed end as each actin subunit is added (1, 9-12). This processive elongation of a barbed end by a formin is terminated when the formin dissociates stochastically from the growing end during translocation (4, 10).The formin-homology (FH)2 1 and 2 domains are the best conserved domains of formin proteins (2, 13, 14). The FH2 domain is the signature domain of formins, and in many cases, is sufficient for both nucleation and processive elongation of barbed ends (2-4, 7, 15). Head-to-tail homodimers of FH2 domains (12, 16) encircle the barbed ends of actin filaments (9). In vitro, association of barbed ends with FH2 domains slows elongation by limiting addition of free actin monomers. This “gating” behavior is usually explained by a rapid equilibrium of the FH2-associated end between an open state competent for actin monomer association and a closed state that blocks monomer binding (4, 9, 17).Proline-rich FH1 domains located N-terminal to FH2 domains are required for profilin to stimulate formin-mediated elongation. Individual tracks of polyproline in FH1 domains bind 1:1 complexes of profilin-actin and transfer the actin directly to the FH2-associated barbed end to increase processive elongation rates (4-6, 8, 10, 17).Rates of elongation and dissociation from growing barbed ends differ widely for FH1FH2 fragments from different formin homologs (4). We understand few aspects of FH1FH2 domains that influence gating, elongation or dissociation. In this study, we examined the source of energy for formin-mediated processive elongation, and the influence of FH2 structure on elongation and dissociation from growing ends. In contrast to previous proposals (6, 18), we found that fast processive elongation mediated by FH1FH2-formins is not driven by energy from the release of the γ-phosphate from ATP-actin filaments. Instead, the data show that the binding of an actin subunit to the barbed end provides the energy for processive elongation. We found that in similar polymerizing conditions, different natural FH2 domains dissociate from growing barbed ends at substantially different rates. We further observed that the length of the flexible linker between the subunits of a FH2 dimer influences dissociation much more than elongation.  相似文献   

20.
Galectins have been implicated in T cell homeostasis playing complementary pro-apoptotic roles. Here we show that galectin-8 (Gal-8) is a potent pro-apoptotic agent in Jurkat T cells inducing a complex phospholipase D/phosphatidic acid signaling pathway that has not been reported for any galectin before. Gal-8 increases phosphatidic signaling, which enhances the activity of both ERK1/2 and type 4 phosphodiesterases (PDE4), with a subsequent decrease in basal protein kinase A activity. Strikingly, rolipram inhibition of PDE4 decreases ERK1/2 activity. Thus Gal-8-induced PDE4 activation releases a negative influence of cAMP/protein kinase A on ERK1/2. The resulting strong ERK1/2 activation leads to expression of the death factor Fas ligand and caspase-mediated apoptosis. Several conditions that decrease ERK1/2 activity also decrease apoptosis, such as anti-Fas ligand blocking antibodies. In addition, experiments with freshly isolated human peripheral blood mononuclear cells, previously stimulated with anti-CD3 and anti-CD28, show that Gal-8 is pro-apoptotic on activated T cells, most likely on a subpopulation of them. Anti-Gal-8 autoantibodies from patients with systemic lupus erythematosus block the apoptotic effect of Gal-8. These results implicate Gal-8 as a novel T cell suppressive factor, which can be counterbalanced by function-blocking autoantibodies in autoimmunity.Glycan-binding proteins of the galectin family have been increasingly studied as regulators of the immune response and potential therapeutic agents for autoimmune disorders (1). To date, 15 galectins have been identified and classified according with the structural organization of their distinctive monomeric or dimeric carbohydrate recognition domain for β-galactosides (2, 3). Galectins are secreted by unconventional mechanisms and once outside the cells bind to and cross-link multiple glycoconjugates both at the cell surface and at the extracellular matrix, modulating processes as diverse as cell adhesion, migration, proliferation, differentiation, and apoptosis (410). Several galectins have been involved in T cell homeostasis because of their capability to kill thymocytes, activated T cells, and T cell lines (1116). Pro-apoptotic galectins might contribute to shape the T cell repertoire in the thymus by negative selection, restrict the immune response by eliminating activated T cells at the periphery (1), and help cancer cells to escape the immune system by eliminating cancer-infiltrating T cells (17). They have also a promising therapeutic potential to eliminate abnormally activated T cells and inflammatory cells (1). Studies on the mostly explored galectins, Gal-1, -3, and -9 (14, 15, 1820), as well as in Gal-2 (13), suggest immunosuppressive complementary roles inducing different pathways to apoptosis. Galectin-8 (Gal-8)4 is one of the most widely expressed galectins in human tissues (21, 22) and cancerous cells (23, 24). Depending on the cell context and mode of presentation, either as soluble stimulus or extracellular matrix, Gal-8 can promote cell adhesion, spreading, growth, and apoptosis (6, 7, 9, 10, 22, 25). Its role has been mostly studied in relation to tumor malignancy (23, 24). However, there is some evidence regarding a role for Gal-8 in T cell homeostasis and autoimmune or inflammatory disorders. For instance, the intrathymic expression and pro-apoptotic effect of Gal-8 upon CD4highCD8high thymocytes suggest a role for Gal-8 in shaping the T cell repertoire (16). Gal-8 could also modulate the inflammatory function of neutrophils (26), Moreover Gal-8-blocking agents have been detected in chronic autoimmune disorders (10, 27, 28). In rheumatoid arthritis, Gal-8 has an anti-inflammatory action, promoting apoptosis of synovial fluid cells, but can be counteracted by a specific rheumatoid version of CD44 (CD44vRA) (27). In systemic lupus erythematosus (SLE), a prototypic autoimmune disease, we recently described function-blocking autoantibodies against Gal-8 (10, 28). Thus it is important to define the role of Gal-8 and the influence of anti-Gal-8 autoantibodies in immune cells.In Jurkat T cells, we previously reported that Gal-8 interacts with specific integrins, such as α1β1, α3β1, and α5β1 but not α4β1, and as a matrix protein promotes cell adhesion and asymmetric spreading through activation of the extracellular signal-regulated kinases 1 and 2 (ERK1/2) (10). These early effects occur within 5–30 min. However, ERK1/2 signaling supports long term processes such as T cell survival or death, depending on the moment of the immune response. During T cell activation, ERK1/2 contributes to enhance the expression of interleukin-2 (IL-2) required for T cell clonal expansion (29). It also supports T cell survival against pro-apoptotic Fas ligand (FasL) produced by themselves and by other previously activated T cells (30, 31). Later on, ERK1/2 is required for activation-induced cell death, which controls the extension of the immune response by eliminating recently activated and restimulated T cells (32, 33). In activation-induced cell death, ERK1/2 signaling contributes to enhance the expression of FasL and its receptor Fas/CD95 (32, 33), which constitute a preponderant pro-apoptotic system in T cells (34). Here, we ask whether Gal-8 is able to modulate the intensity of ERK1/2 signaling enough to participate in long term processes involved in T cell homeostasis.The functional integration of ERK1/2 and PKA signaling (35) deserves special attention. cAMP/PKA signaling plays an immunosuppressive role in T cells (36) and is altered in SLE (37). Phosphodiesterases (PDEs) that degrade cAMP release the immunosuppressive action of cAMP/PKA during T cell activation (38, 39). PKA has been described to control the activity of ERK1/2 either positively or negatively in different cells and processes (35). A little explored integration among ERK1/2 and PKA occurs via phosphatidic acid (PA) and PDE signaling. Several stimuli activate phospholipase D (PLD) that hydrolyzes phosphatidylcholine into PA and choline. Such PLD-generated PA plays roles in signaling interacting with a variety of targeting proteins that bear PA-binding domains (40). In this way PA recruits Raf-1 to the plasma membrane (41). It is also converted by phosphatidic acid phosphohydrolase (PAP) activity into diacylglycerol (DAG), which among other functions, recruits and activates the GTPase Ras (42). Both Ras and Raf-1 are upstream elements of the ERK1/2 activation pathway (43). In addition, PA binds to and activates PDEs of the type 4 subfamily (PDE4s) leading to decreased cAMP levels and PKA down-regulation (44). The regulation and role of PA-mediated control of ERK1/2 and PKA remain relatively unknown in T cell homeostasis, because it is also unknown whether galectins stimulate the PLD/PA pathway.Here we found that Gal-8 induces apoptosis in Jurkat T cells by triggering cross-talk between PKA and ERK1/2 pathways mediated by PLD-generated PA. Our results for the first time show that a galectin increases the PA levels, down-regulates the cAMP/PKA system by enhancing rolipram-sensitive PDE activity, and induces an ERK1/2-dependent expression of the pro-apoptotic factor FasL. The enhanced PDE activity induced by Gal-8 is required for the activation of ERK1/2 that finally leads to apoptosis. Gal-8 also induces apoptosis in human peripheral blood mononuclear cells (PBMC), especially after activating T cells with anti-CD3/CD28. Therefore, Gal-8 shares with other galectins the property of killing activated T cells contributing to the T cell homeostasis. The pathway involves a particularly integrated signaling context, engaging PLD/PA, cAMP/PKA, and ERK1/2, which so far has not been reported for galectins. The pro-apoptotic function of Gal-8 also seems to be unique in its susceptibility to inhibition by anti-Gal-8 autoantibodies.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号